A Crossed Molecular Beam Study of the Reaction ... - ACS Publications

In previous work from this laboratory we reported the results of direct dynamical investigations, using the crossed molecular beam (CMB) scattering me...
0 downloads 0 Views 277KB Size
J. Phys. Chem. A 1997, 101, 6455-6462

6455

A Crossed Molecular Beam Study of the Reaction O(1D) + HI f IO + H Michele Alagia, Nadia Balucani, Piergiorgio Casavecchia,* and Gian Gualberto Volpi Dipartimento di Chimica, UniVersita` di Perugia, 06123 Perugia, Italy ReceiVed: February 25, 1997; In Final Form: May 19, 1997X

The angular and velocity distributions of the IO product from the reaction O(1D) + HI have been obtained in crossed beam experiments at collision energies, Ec, of 4.7 and 13.6 kcal/mol. The product center-of-mass angular distribution is found to be nearly backward-forward symmetric, with forward scattering favored. The forward to backward scattering ratio increases with increasing Ec, from which it is deduced that the reaction occurs via an intermediate complex having a lifetime comparable to its rotational period. Within the “osculating” model for chemical reactions, the complex lifetime is estimated to be 4 and 1 ps at low and high Ec, respectively. A large fraction (0.55 at low Ec and 0.46 at high Ec) of the total available energy was found to be released into product translational energy. Comparison with the dynamics of the analogous reactions O(1D) + HCl and O(1D) + HBr is carried out: similarities as well as differences are noted and discussed in terms of the potential energy surfaces for reaction. From comparison of signal intensities and the previously estimated lower limits to the branching ratio of the relative cross sections for XO + H and OH + X channels in the O(1D) + HX (X ) Cl, Br) reactions, it is concluded that the H-displacement pathway accounts for a significant fraction of the overall reaction also in the case of O(1D) + HI.

Introduction The reactions of atomic oxygen, in both the ground state O(3P) and the first electronically excited state O(1D), with halogenated compounds are of interest in determining the impact of anthropogenic surface release of halogen-containing molecules on the atmosphere and especially on the ozone natural balance. They are also of relevance in the combustion chemistry of halogenated compounds, as for instance in their use as fire extinguishers. Among these reactions, those with hydrogen halides are particularly simple and can be considered as prototypes; indeed, they comprise a class of three-atom reactions that are amenable to detailed experimental and theoretical investigation. While the reactions of HCl and HBr with both O(3P) and O(1D) have been recently studied from both kinetics1-4 and dynamics5-9 standpoints, little information is available on the reaction of HI with O(3P)3,10,11 and none on the reaction with O(1D). In previous work from this laboratory we reported the results of direct dynamical investigations, using the crossed molecular beam (CMB) scattering method, on the reactions O(1D) + HCl,5 HBr,7 and CF3Br,12 relative to the XO (X ) Cl, Br) formation channel. Here we extend these studies to the next member of the O(1D) + HX series, namely, the reaction O(1D) + HI. This reaction can proceed via two competing pathways:

O(1D) + HI f IO + H f OH + I

∆H°0 ≈ -30 kcal/mol

(1a)

∆H°0 ) -75.8 kcal/mol (1b)

Heats of reaction were derived from the data of ref 13, with the exception of ∆H°f,0 (IO) (see below). All products of reaction 1 may be catalytically active species in ozone chemistry. Recently, a growing concern for the potential role of iodine in atmospheric chemistry has motivated experimental and theoretical studies of the likely relevant iodine “active” and “reservoir” forms and iodine involving reactions. In 1980, Chameides and Davis14 first pointed out that, while the active forms of chlorine X

Abstract published in AdVance ACS Abstracts, July 15, 1997.

S1089-5639(97)00697-X CCC: $14.00

and bromine contribute to stratospheric ozone depletion, the tropospheric ozone balance may be affected significantly by iodine compounds. In fact, in the tropospherical environment the thermal and photochemical stability of the principal inorganic reservoirs of chlorine and bromine (i.e., HCl, HBr, HOCl, HOBr, ClONO2, and BrONO2) greatly limits the impact of these halogens on the ozone budget. On the contrary, iodine reservoir species are easily photodissociated in the near-UV and visible region (since the bonds involving iodine are generally weak), leading to the rapid release of active iodine; the dominant reaction of the formed iodine atoms seems to be that with O3

I + O3 f IO + O2

(2)

while the IO radical returns back to I primarily by photodissociation

IO + hν (λ ) 4550 Å) f I + O

(3)

IO + NO f I + NO2

(4)

or by reaction

The IO radical can also react to form reservoir species through

IO + NO2 f IONO2 IO + HO2 f HOI + O2 f HI + O3

(5) (6a) (6b)

Of all the inorganic reservoirs, HI is the least photolabile (its photochemical lifetime is ≈1-2 h during the day), and its reaction with OH radicals

HI + OH f H2O + I

(7)

is the main gas-phase HI loss process. Reaction 7 regenerates the active form of I. Many other reactions, including photodissociation of HOI, were postulated to occur, giving rise to a net loss of tropospheric ozone. Since most of these reactions were not studied at that time, the potential role of iodine was © 1997 American Chemical Society

6456 J. Phys. Chem. A, Vol. 101, No. 36, 1997 just proposed by the authors,14 who made a call for experimental studies and field measurements in order to assess the model. Subsequently, most of the reactions involved and the photochemical processes proposed were studied and rate coefficients determined.15 Lately, there has been a renewed interest in iodine chemistry stemming from the recent assessment that iodine could account for the observed depletion of lower stratospheric ozone (below 20 km, altitudes where chlorine and bromine are not very effective for ozone destruction in contrast to iodine) at midlatitudes.16 Studies of convective transport have shown that vertical transport of iodine-containing molecules can be significant, allowing the photochemically short-lived iodine compounds to reach the upper troposphere and even the lower stratosphere. Due to the instability of its reservoir species, the efficiency of iodine released in the stratosphere for ozone depletion near 15-20 km would be more than 1000 times greater than that of Cl, leading to the conclusion that even a very small amount of iodine could cause significant O3 destruction.16 Actually, the iodine role in ozone depletion is thought to be of enhancing the effect of the human-released chlorine and bromine, through the interhalogen reactions IO + ClO and IO + BrO.16 The main source of atmospheric iodine is natural: methyl iodide, chloroiodomethane, and diiodomethane are thought to be produced by marine biota, such as phytoplankton, kelp, and macroalgae; in particular, iodocarbons generated by biomass burning could contribute significantly to the iodine reaching directly the stratosphere.16 The fraction of stratospheric iodine that is expected to reside in the form of HI is ≈0.5%.16 Even though reaction (1) is not thought to play a major role either in the tropospheric ozone budget (where reaction 7 is dominant) or in the stratospheric one, in both environments O(1D) is significantly present since it is the product of ozone photodissociation. All the products of reaction 1 are active forms in ozone depletion, so that reaction 1 could represent an alternative way for iodine to pass from the reservoir HI to active radicals. Moreover, reaction 1 can be considered as prototypical of the reactions of O(1D) with simple iodine-containing molecules, a class of compounds which is now being considered as a substitute for halon fire suppressants.16b The rate constant of reaction 1 has never been determined; recent kinetic studies provided room-temperature rate constants for the analogous reactions O(1D) + HCl and O(1D) + HBr (and also O(1D) + HF).4a Comparison with the rate constants for the reactions O(3P) + HCl and O(3P) + HBr shows that the electronic excitation of the oxygen atom leads to an increase of 6 and 4 orders of magnitude for HCl and HBr, respectively. For O(3P) + HI f OH + I it was found3 k298 ) 1.62 × 10-12 cm3 molecule-1 s-1, quite larger than 3.8 × 10-14 and 1.3 × 10-16 cm3 molecule-1 s-1, which are the k298 values for O(3P) + HBr and O(3P) + HCl, respectively.1-3 Therefore, the rate constant for reaction 1 is expected to be at least comparable to those of the reactions O(1D) + HCl and O(1D) + HBr, especially if one considers that the reactions of O(1D) with most compounds usually occur with near gas kinetic frequency due to the fact that they take place on a singlet ground state potential energy surface (PES) with virtually zero activation energy. It should be noted that for the O(3P) + HX (X ) Cl, Br) reactions the OH + X channel is exoergic and the H + XO channel strongly endoergic. Instead, both channels are exoergic for O(1D); in particular, kinetic4 and dynamic5,6d,7,9a studies indicate that the H-displacement channel in the O(1D) + HX (X ) Cl, Br) reactions accounts for a considerable fraction of the overall reaction. Dynamical studies of O(1D) reactions with simple molecules also address the important issue of the role of electronic

Alagia et al. excitation on the reaction dynamics of nonmetal atoms. The reactions of O(1D) with H-X molecules (where X ) H, Cl, Br, I, OH, SH, CH3) are postulated to proceed through the formation of a strongly bound complex, HOX, following the insertion of O(1D) into a bond of the colliding molecule.17 However, the lifetimes of the complexes and the dynamics of their dissociation are not well-known yet. In particular, the nonstatistical energy partitioning observed in most of the O(1D) reactions pointed to a precluded equilibration of the available energy among the degrees of freedom of the complexes;6f in other words, the randomization of energy of the complex does not occur within the time scale required for its decomposition. For instance, in the case of the reaction O(1D) + CH4, even though the decay time of the CH3OH formed from the photolysis of the CH4‚O3 van der Waals adduct was found18 to be 3 ps at Ec ≈ 13 kcal/mol (a value comparable to the lifetime calculated from Rice-Ramsperger-Kassel-Marcus theory for the dissociation of methanol at the level of excitation of the experiment), the OH product internal energy distribution is not found to be statistical.19 The results presented below, together with the information obtained in this laboratory on the related reactions O(1D) + H2,20 O(1D) + HCl,5 O(1D) + HBr,7 O(3P,1D) + H2S,21 and O(3P,1D) + CH3I,22 provide a useful database also for comparison with theory. It should be noted that reactions involving O(3P,1D) with simple molecules are good candidates for studies investigating the role of multiple potential energy surfaces and nonadiabatic effects on chemical reaction dynamics. The study reported here has been carried out using the powerful CMB method pioneered by Professor Y. T. Lee, together with Professor D. R. Herschbach.23 Also, the method used for generating intense continuous supersonic beams of O(1D) atoms, which is crucial for the present studies, has been originally developed in Lee’s laboratory.24 It is our pleasure and most appropriate to dedicate this paper to Professor Lee on the occasion of his 60th birthday. Experimental Section The scattering experiments were carried out on the universal crossed molecular beam apparatus of Perugia, which is an implementation for reactive scattering studies of the highresolution machine previously used for measuring elastic and total (elastic + inelastic) differential cross sections for atomatom and atom-molecule collisions.25 The apparatus has been described in detail elsewhere,25,26 and only a brief description necessary to understand the present experiments is given here. Two well-collimated supersonic beams of the reactants, after two stages of differential pumping, are crossed at 90° under single-collision conditions in a large scattering chamber maintained below 5 × 10-7 mbar. The reaction products are detected by an electron bombardment ionization quadrupole mass spectrometer which can rotate in the plane of the two beams around their intersection axis (i.e., around the collision region). The ionizer is mounted on a liquid nitrogen cooled OFHC copper chamber, which is kept in the 10-11 mbar region by extensive ion-, turbo-, and cryo-pumping and represents the innermost region of a triply differentially pumped ultrahighvacuum rotatable chamber. The seeded supersonic atomic oxygen beams, containing O(1D) and ground-state O(3P), were generated by a highpressure, radio-frequency (rf) discharge nozzle source similar in design to that described by Sibener et al.24 and which has been optimized for O(1D) production.5,7,20,21,26 Two O(1D) beams with different translational energies were produced in order to vary the relative collision energy of the experiment.

O(1D) + HI f IO + H By discharging 350 mbar of a 5% O2 in helium gas mixture through a 0.26 mm diameter quartz nozzle at 250 W of rf power, a beam with peak velocity of 2854 m/s and speed ratio of 7.0 was obtained. By discharging 350 mbar of a 5% O2 in neon gas mixture through the same quartz nozzle at 290 W of rf power, a beam with peak velocity of 1661 m/s and speed ratio of 6.2 was produced. The angular divergence is 2.3° in both cases. Using the fast and slow beams of O(1D), relative collision energies, Ec, of 13.6 and 4.7 kcal/mol, respectively, were obtained (see below). Due to the very efficient coupling of the rf power to the plasma, high degrees of molecular dissociation are achieved (>80% and >90% with the helium and neon mixture, respectively). The oxygen beams contain a very large fraction of ground-state O(3P), but this species does not interfere since the collision energies of the present experiments are not sufficient to overcome the large endoergicity of the reaction O(3P) + HI f IO + H (∆H°r ≈ 15.4 kcal/mol). Supersonic beams of HI were generated by expanding pure HI through a 100 µm diameter stainless steel nozzle. For the experiment at Ec ) 13.6 kcal/mol the HI stagnation pressure was 0.35 bar, and the resulting peak velocity and speed ratio were 370 m/s and 8.1, respectively; for the experiment at Ec ) 4.7 kcal/mol slightly different experimental conditions, such as the HI stagnation pressure which was 0.32 bar, gave a peak velocity of 369 m/s and a speed ratio of 8.0. The angular divergence was 2.9° in both cases. Under these expansion conditions HI clustering was negligible. Because of the significant rotational cooling during supersonic expansion (a rotational temperature of about 16 K is estimated, assuming equilibrium between translational and rotational degrees of freedom), the HI molecules in the beam are expected to be in the lowest few rotational states of the ground vibrational level, and therefore the internal energy of the molecular reactant contributes negligibly (> L′ and most of the total angular momentum is taken away from the rotation of the molecules, i.e., j′ = L. Product Energy Distributions. A common feature of the XO (X ) Cl, Br, I) + H reaction channel is the large fraction (about 50%) of the total available energy released into product translation. This fraction is somewhat larger than that predicted by the RRHO microcanonical prior distribution of Levine and Bernstein30 for these systems and may be taken as an indication of the presence of a small potential barrier in the exit valley, as suggested by classical trajectory studies47 on model surfaces for the H + HL f HH + L (H ) heavy, L ) light) mass combination. Interestingly, recent ab initio and model calculations of the PES found a barrier for dissociation of HClO9a,b and also of HBrO;48 it is reasonable to expect an exit barrier for dissociation of HIO as well. This suggests that HIO complexes actually play a role in the O(1D) + HI f IO + H reaction. It is interesting to note that both P(E′T) distributions (Figure 3b) peak at a value of about 19 kcal/mol, which is significantly lower than the exoergicity of the reaction (∆Hr ) -30 kcal/ mol), while for the reactions with HCl and HBr the correspond-

O(1D) + HI f IO + H ing P(E′T) peaked beyond the reaction exoergicity. However, while for O(1D) + HBr the 〈E′T〉 fraction did not vary with Ec, for O(1D) + HI it decreases considerably with increasing Ec (from 55% to 46%), which indicates that the extra initial translational energy is more efficiently channeled into internal energy. Since a very large fraction of the available energy is released as product translation and high rotational excitation also occurs (assuming j′ ) L, it can be estimated an average product rotational energy 〈E′R〉 of 5.3 and 10.6 kcal/mol at low and high Ec, respectively), little energy remains for electronic and vibrational excitation of the IO product. Since the spinorbit splitting of IO is 6 kcal/mol and the vibrational spacing about 2 kcal/mol, IO should be mainly formed in the lowest few vibrational levels of both possible 2Π1/2 and 2Π3/2 states. Unfortunately, in the present experiment the relative population of the fine structure levels of IO could not be disentangled during the data analysis. Branching Ratio σ(IO + H)/σ(OH + I). A lower limit for the ratio between the cross section, σ, for formation of XO and OH for the analogous O(1D) + HCl5a and O(1D) + HBr7 reactions was estimated to be σ(ClO + H)/σ(OH + Cl) ) 0.34 ( 0.10 and σ(BrO + H)/σ(OH + Br) ) 0.16 ( 0.07. For X ) I no estimate was attempted, because the sensitivity to the OH formation channel decreases as the halogen atom mass increases, due to the very unfavorable kinematics. Kinematically, it should be more convenient to detect the I atom than the OH counterpart, but unfortunately, I + produced from dissociative ionization, under electron impact in the detector, of elastically scattered HI represents a severe complication. However, on the basis of the relative signal intensities observed (S(IO) ≈ 2.5S(BrO) ≈ 7S(ClO) under approximately the same experimental conditions), it is inferred that the IO formation channel should be as important as the BrO and ClO formation channels and, therefore, a significant fraction of the corresponding overall O(1D) + HI reaction. The findings for X ) Cl, Br were in agreement with bulk estimates.4 Conclusion The dynamics of the O(1D) + HI reaction relative to the H-displacement channel is found to resemble those of the analogous O(1D) + HCl and O(1D) + HBr reactions, but with the noteworthy difference that no evidence for a direct rebound mechanism arising from collinear attack of the oxygen atom on the halogen side of the hydride is observed in this case. IO formation is found to occur via an osculating complex, whose lifetime decreases with increasing collision energy. Complex formation arises likely from an insertion mechanism involving the ground state PES X1A′ of HOI and/or from an addition mechanism leading to HIO, which may subsequently isomerize to HOI or directly dissociate to IO + H. Interestingly, the dynamics of the reaction O(1D) + HI f IO + H are very similar to those of O(1D) + H2S f HSO + H, where the absence of a backward scattered component was also noted and the complex lifetime was decreasing with increasing Ec. No information is available on the OH forming channel. In the two other reactions of the series, spectroscopic studies6,7 found OH to be formed highly vibrationally and rotationally excited. The explanation proposed was that the reactions proceed through the insertion of O(1D) into the H-X bond on the 1A′ surface with formation of an internally hot HOX complex, which fragments within a few bending vibrations into OH + X, before the randomization of energy is complete. The OH forming channel in all the O(1D) reactions studied to date by using infrared chemiluminescence or laser-induced fluorescence has been found to have a nonstatistical product internal

J. Phys. Chem. A, Vol. 101, No. 36, 1997 6461 energy distribution, indicating that complete energy randomization does not occur within the complex lifetime; however, this does not necessarily indicate that the complex lifetime is short in comparison to its average rotational period.6f,19d Our results on the H-displacement channel examined in this and in other related systems indicate that the reactions proceed mainly through the formation of a collision complex living a few rotational periods. When X changes along the series Cl, Br, and I, the two possible XO and OH forming channels become more exoergic, and the complex lifetime decreases with respect to its rotational period. The present experimental results complement the dynamical information obtained on the other members of the O(1D) + HX series and should facilitate attempts to model the dynamics of these reactions using realistic PESs. An accurate description of the fate of the presumably formed HOX/HXO complexes will be possibly given by quasiclassical trajectory and quantum mechanical dynamics calculations when they will become available. Theoretical work along these lines has recently appeared in the literature for the O(1D) + HCl reaction,9a,b is in progress for O(1D) + HBr,48 and is being planned for O(1D) + HI.49 Acknowledgment. We acknowledge useful discussions with Prof. J. M. Farrar, Prof. J. J. Sloan, and Prof. A. Lagana`. We thank Prof. P. Marshall for communicating to us his results prior to publication. Also, we thank an anonymous referee for useful comments and suggestions. This work has been supported by the Italian “Consiglio Nazionale delle RicerchesProgetto Finalizzato Chimica Fine” and “Ministero Universita` e Ricerca Scientifica”. Grants from NATO (CRG 920549) and AFOSR (F617-08-94-C-0013) through EOARD (SPC-94-4042) made possible several aspects of the research reported here. References and Notes (1) (a) Ravishankara, A. R.; Smith, G.; Watson, R. T.; Davis, D. D. J. Phys. Chem. 1977, 81, 2220. (b) Mahmud, K.; Kim, J. S.; Fontijn, A. J. Phys. Chem. 1990, 94, 2994. (2) (a) Brown, R. D. H.; Smith, I. W. M. Int. J. Chem. Kinet. 1975, 7, 301; 1978, 10, 1. (b) Nava, D. F.; Bosco, S. R.; Stief, L. J. J. Chem. Phys. 1983, 78, 2443. (3) Singleton, D. L.; Cvetanovic, R. J. Can. J. Chem. 1978, 56, 2934. (4) (a) Wine, P. H.; Wells, J. R.; Ravishankara, A. R. J. Chem. Phys. 1986, 84, 1349. (b) Wine, P. H. Private communication. (5) (a) Balucani, N.; Beneventi, L.; Casavecchia, P.; Volpi, G. G. Chem. Phys. Lett. 1991, 180, 34. (b) Balucani, N.; Beneventi, L.; Casavecchia, P.; Stranges, D.; Volpi, G. G. In Ecological Physical Chemistry; Rossi, C.; Tiezzi, E., Eds.; Elsevier: Amsterdam, 1991; p 539. (6) (a) Luntz, A. C. J. Chem. Phys. 1980, 73, 5393. (b) Kruus, E. J.; Niefer, B. I.; Sloan, J. J. J. Chem. Phys. 1988, 88, 985. (c) Park, C. R.; Wiesenfeld, J. R. Chem. Phys. Lett. 1989, 163, 230. (d) Matsumi, Y.; Tarokura, K.; Kawasaki, M.; Tsuji, K.; Obi, K. J. Chem. Phys. 1993, 98, 8330. (e) Matsumi, Y.; Shamsuddin, S. M. J. Chem. Phys. 1995, 103, 4490. (f) Alexander, A. J.; Brouard, M.; Rayner, S. P.; Simons, J. P. Chem. Phys. 1996, 207, 215. (7) Balucani, N.; Beneventi, L.; Casavecchia, P.; Volpi, G. G.; Kruus, E. J.; Sloan, J. J. Can. J. Chem. 1994, 72, 888. (8) (a) McKendrick, K. G.; Rakestraw, D. J.; Zhang, R; Zare, R. N. J. Phys. Chem. 1988, 92, 5530. (b) Zhang, R.; van der Zande, W. J.; Bronikowski, M. J.; Zare, R. N. J. Chem. Phys. 1991, 94, 2704. (9) (a) Lagana`, A.; Ochoa de Aspuru, G.; Garcia, E. J. Phys. Chem. 1995, 99, 17139. (b) Hernandez, M. L.; Redondo, C.; Lagana`, A.; Ochoa de Aspuru, G.; Rosi, M.; Sgamellotti, A. J. Chem. Phys. 1996, 105, 2710. (c) Schinke, R. J. Chem. Phys. 1984, 80, 5510. (d) Koizumi, H.; Schatz, G. C.; Gordon, M. S. J. Chem. Phys. 1991, 95, 6421. (10) Agrawalla, B. S.; Manocha, A. S.; Setser, D. W. J. Phys. Chem. 1981, 85, 2873. (11) Persky, A.; Broida, M. Chem. Phys. 1987, 114, 85. (12) Alagia, M.; Balucani, N.; Casavecchia, P.; Lagana`, A.; Ochoa de Aspuru, G.; van Kleef, E. H.; Volpi, G. G.; Lendvay, G. Chem. Phys. Lett. 1996, 258, 323. (13) Atkinson, R.; Baulch, D. L.; Cox, R. A.; Hampson, R. F., Jr.; Kerr, J. A.; Troe, J. J. Phys. Chem. Ref. Data 1992, 21, 1125. (14) Chameides, W. L.; Davis, D. D. J. Geophys. Res. 1980, 85, 7383.

6462 J. Phys. Chem. A, Vol. 101, No. 36, 1997 (15) (a) Jenkin, M. E.; Cox, R. A. J. Phys. Chem. 1985, 89, 192. (b) Sander, S. P. J. Phys. Chem. 1986, 90, 2194. (c) Stickel, R. E.; Hynes, A. J.; Bradshaw, J. D.; Chameides, W. L.; Davis, D. D. J. Phys. Chem. 1988, 92, 1862. (d) Daykin, E. P.; Wine, P. H. J. Phys. Chem. 1990, 94, 4528. (e) Mac Leod, H.; Balestra, C.; Jourdain, J. L.; Laverdet, G.; Le Bras, G. Int. J. Chem. Kinet. 1990, 22, 1167. (f) Maguin, F.; Laverdet, G.; Le Bras, G.; Poulet, G. J. Phys. Chem. 1992, 96, 1775. (g) Turnipseed, A. A.; Gilles, M. K.; Burkholder, J. B.; Ravishankara, A. R. Chem. Phys. Lett. 1995, 242, 427. (h) Bedjanian, Y.; Le Bras, G.; Poulet, G. J. Phys. Chem. 1996, 100, 15130. (i) Harwood, M. H.; Burkholder, J. B.; Hunter, M.; Fox, R. W.; Ravishankara, A. R. J. Phys. Chem. A 1997, 101, 853. (16) (a) Solomon, S.; Garcia, R. R.; Ravishankara, A. R. J. Geophys. Res. 1994, 99 (D10), 20491. See also: (b) Solomon, S.; Burkholder, J. B.; Ravishankara, A. R.; Garcia, R. R. J. Geophys. Res. 1994, 99 (D10), 20929. (17) Sloan, J. J. J. Phys. Chem. 1988, 92, 18. (18) van Zee, R. D.; Stephenson, J. C. J. Chem. Phys. 1995, 102, 6946. (19) (a) Luntz, A. C. J. Chem. Phys. 1980, 73, 1143. (b) Aker, P. M.; O’Brien, J. J. A.; Sloan, J. J. J. Chem. Phys. 1986, 84, 745. (c) Park, C. R.; Wiesenfeld, J. R. J. Chem. Phys. 1991, 95, 8166. (d) Brouard, M.; Duxon, S.; Enriquez, P. A.; Simons, J. P. J. Chem. Soc., Faraday Trans. 1993, 89, 1435. (20) Alagia, M.; Balucani, N.; Cartechini, L.; Casavecchia, P.; G.; van Kleef, E. H.; Volpi, G. G.; Kruus, P. J.; Sloan, J. J. Manuscript in preparation. (21) (a) Balucani, N.; Beneventi, L.; Casavecchia, P.; Stranges, D.; Volpi, G. G. J. Chem. Phys. 1991, 94, 8611. (b) Balucani, N.; Casavecchia, P.; Stranges, D.; Volpi, G. G. Chem. Phys. Lett. 1993, 211, 469. (c) Balucani, N.; Casavecchia, P.; Stranges, D.; Volpi, G. G. Manuscript in preparation. (22) Alagia, M.; Balucani, N.; Cartechini, L.; Casavecchia, P.; van Beek, M. C.; Volpi, G. G., Manuscript in preparation. (23) Lee, Y. T.; McDonald, J. D.; LeBreton, P. R.; Herschbach, D. R. ReV. Sci. Instrum. 1969, 40, 1402. (24) Sibener, S. J.; Buss, R. J.; Ng, C. Y.; Lee, Y. T. ReV. Sci. Instrum. 1980, 51, 167. (25) (a) Beneventi, L.; Casavecchia, P.; Volpi, G. G. J. Chem. Phys. 1986, 85, 7011. (b) Beneventi, L.; Casavecchia, P.; Volpi, G. G., Bieler, C. R.; Janda, K. C. J. Chem. Phys. 1993, 98, 178 and references therein. (26) (a) Casavecchia, P.; Balucani, N.; Volpi, G. G. In Research in Chemical Kinetics; Compton, R. G., Hancock, G., Eds.; Elsevier: Amsterdam, 1993; Vol. 1, p 1. (b) Casavecchia, P.; Balucani, N.; Volpi, G. G. In AdVanced Series in Physical Chemistry; Wagner, A., Liu, K., Eds.; Worlds Scientific: Singapore, 1995; Vol. 6, p 365. (c) Alagia, M.; Balucani, N.; Casavecchia, P.; Stranges, D.; Volpi, G. G. J. Chem. Soc., Faraday Trans. 1995, 91, 575. (27) Sko¨ld, K. Nucl. Instrum. Methods 1968, 63, 114.

Alagia et al. (28) Buss, R. J. Ph.D. Thesis, University of California, Berkeley, 1979. (29) Lee, Y. T. In Atomic and Molecular Beam Methods: Scoles, G., Ed.; Oxford Univesity Press: New York, 1987; Vol. 1. (30) Levine, R. D.; Bernstein, R. B. Molecular Reaction Dynamics; Oxford Univesity Press: Oxford, 1974. (31) Chase, M. W. J. Phys. Chem. Ref. Data, 1996, 25, 1297 and references therein. (32) Coleman, E. H.; Gaydon, A. G.; Vaidya, W. M. Nature 1948, 162, 108. (33) Radlein, D. St. A. G.; Whitehead, J. C.; Grice, R. Nature 1975, 253, 5486. (34) Buss, R. J.; Sibener, S. J.; Lee, Y. T. J. Phys. Chem. 1983, 87, 4840. (35) Gilles, M. K.; Turnipseed, A. A.; Rudich, Y.; Talukdar, R. K.; Villalta, P.; Huey, L. G.; Burkholder, L. B.; Ravishankara, A. R. J. Phys. Chem. 1996, 100, 14005. (36) McGrath, M. P.; Rowland, F. S. J. Phys. Chem. 1996, 100, 4815. (37) Hassanzadeh, P. H.; Irakura, K. K.; Johnson, R. D. III Unpublished data; see ref 31. (38) (a) Bekoy, J. P.; Leo Meerts, W.; Dymanus, A. J. Mol. Spectrosc. 1983, 102, 320. (b) Gilles, M. K.; Polak, M. L.; Lineberger, W. C. J. Chem. Phys. 1991, 95, 4723; 1992, 96, 8012. (39) About the enthalpy of formation of HOI we refer to the value reported in; Loewenstein, L. M.; Anderson, J. G. J. Phys. Chem. 1987, 91, 2993. (40) (a) Klaassen, J. J.; Lindner, J.; Leone, S. R. J. Chem. Phys. 1996, 104, 7403. (b) Barnes, I.; Becker, K. H.; Starke, J. Chem. Phys. Lett. 1992, 196, 578. (c) Jenkin, M. E.; Clemitshaw, K. C.; Cox, R. A. J. Chem. Soc., Faraday Trans. 2 1984, 80, 1633. (41) Walker, N.; Tevault, D. E.; Smardzewski, R. R. J. Chem. Phys. 1978, 69, 564. (42) Pasquette, J.; Ford, B. L. Can. J. Chem. 1985, 63, 2444. (43) Bruna, P. J.; Horsch, G.; Peyerimhoff, S. D.; Buenker, R. J. Can. J. Chem. 1979, 57, 1839. (44) Marshall, P. Private communication. (45) Miller, W. B.; Safron, S. A.; Herschbach, D. R. Discuss. Faraday Soc. 1967, 44, 108. (46) Fisk, G. A.; McDonald, J. D., Herschbach, D. R. Discuss. Faraday Soc. 1967, 44, 228. (47) Polanyi, J. C.; Schreiber, J. L. In Physical Chemistry: An AdVanced Treatise, Jost, W., Ed.; Academic Press: New York, 1974; Vol. 6A, p 383. (48) Ochoa de Aspuru, G. Ph.D. Thesis, Universidad del Pais Vasco, 1996. Lagana`, A.; Ochoa de Aspuru, G.; Garcia, E. Manuscript in preparation. (49) Lagana`, A. Private communication.